You are on page 1of 11

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/361240919

Experimental study of cavitation index in an ogee spillway by considering


convergence angle of sidewalls

Article  in  Water Supply · June 2022


DOI: 10.2166/ws.2022.228

CITATION READS

1 117

2 authors, including:

Reza Barati
Water Authority
125 PUBLICATIONS   1,350 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Water resources measurement using data loggers (Remote monitoring) View project

Ogee stepped spillway View project

All content following this page was uploaded by Reza Barati on 18 June 2022.

The user has requested enhancement of the downloaded file.


Uncorrected Proof

© 2022 The Authors Water Supply Vol 00 No 0, 1 doi: 10.2166/ws.2022.228

Experimental study of cavitation index in an ogee spillway by considering convergence


angle of sidewalls

Ali Foroudia,* and Reza Barati b


a
Civil Engineering Department, Quchan University of Technology, Quchan, Iran
b
Department of Civil Engineering, Tarbiat Modares University, Tehran, Iran
*Corresponding author. E-mail: aliforoudi@qiet.ac.ir

RB, 0000-0003-2362-2227

ABSTRACT

Cavitation is among the most complex and common damages occurring in spillway structures, which is one of the most expensive parts of a
dam. The cavitation index, as the one of the most efficient approaches, can be used to analyze this important hydraulic phenomenon. The
present study examines the changes in the cavitation index caused by changes in the convergence angle of an ogee spillway’s sidewalls with
an arc in the plan. To this end, a spillway was constructed on the 1:50 scale. Then, it was tested with four different convergence angles,
including 0°, 60°, 90°, and 120°, relative to the spillway’s sidewalls and six different flow rates per unit width ranging from 6.74 to 48.42
(l/s)/m. The results indicated that as the flow rate increased, the cavitation index relatively declined at both crest and chute of the spillway
while growing at its toe. It was observed that the lowest cavitation index was found to be 1.54 in.X/Hd ¼ 2.42 at an angle of 0° and a flow rate
per unit width of 40.52 (L/s)/m.

Key words: guide walls, hydraulic performance, ogee spillway, physical model

HIGHLIGHTS

• Experimental study of convergence angle effects of ogee spillway’s sidewalls on the cavitation index.
• Considering the range of 0° to 120° for convergence angle.
• Considering six different flow rates per unit width in the range of 6.74–48.42 (l/s)/m.
• Presenting and discussing variation of the cavitation index.

This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY 4.0), which permits copying, adaptation and
redistribution, provided the original work is properly cited (http://creativecommons.org/licenses/by/4.0/).

Downloaded from http://iwaponline.com/ws/article-pdf/doi/10.2166/ws.2022.228/1061245/ws2022228.pdf


by guest
Uncorrected Proof

Water Supply Vol 00 No 0, 2

GRAPHICAL ABSTRACT

INTRODUCTION
Chutes and ogee spillways are utilized as the most popular structures in the construction of dams (Wei 2006; Barati 2012;
Shahheydari et al. 2015; Hosseini et al. 2016; Arami-Fadafan et al. 2018a; Bananmah et al. 2020; Tajnesaie et al. 2020).
These types of structures are at the risk of cavitation due to the high level and velocity of water flow (Arami-Fadafan et al.
2018b). The cavitation is defined as the formation of a bubble or void within a liquid. The phenomena associated with cavita-
tion are among the most important factors in which many hydraulic engineers are interested. The cavitation should be
considered in both study and design of dam construction and other similar projects. The cavitation bubbles will collapse
as they travel to areas with relatively higher local pressure (Rafi et al. 2012; Kermani et al. 2013; Rajasekhar et al. 2014;
Usta 2014; Adama Maiga et al. 2016; Koen 2017; Yusuf & Micovic 2020; Karimi Pirmoosaei & Mardookhpour 2020;
Kocaer & Yarar 2020). The cavitation can cause damages to a spillway or create a hole at high speeds of flow current, as
the collapse of vapor cavities results in high pressure shock waves. Ogee spillways are widely used in the design of hydraulic
structures because of their ability to release surplus water from upstream to downstream efficiently and safely when properly
designed and implemented. In order to gain a better understanding of the ogee spillways and their characteristics, it is also
understood that a deviation from the standard design parameters such as a change in upstream flow conditions, modified
crest shape, or change in approach channel owing to local geometric properties can change the flow properties. Therefore,
it is essential to test a hydraulic model with different spillways relevant to various dams with certain geometric conditions
since the construction of dams and relevant facilities is costly. Moreover, there may be possible property damages, as well
as loss of lives, if a spillway does not function properly.
Ball (1976) examined the effects of flow on nonplanar surfaces on cavitation. He showed that cavitation occurs in velocities
higher than 30 m/s, even indents equal to 3 mm. Holand and Venp (1970) were among those who studied this phenomenon
with quick photography. Wang and Chou (1970) performed several studies on the measurement of cavitation index and esti-
mation of damages caused by cavitation. Chanson (1988) conducted comprehensive studies on aeration and aeration devices
in a spillway model. He expressed that cavitation begins in the presence of poor nuclei. Kramer (2004) stated that cavitation
mostly occurs in hydraulic machines and structures. This phenomenon lies in low pressure associated with high speeds,

Downloaded from http://iwaponline.com/ws/article-pdf/doi/10.2166/ws.2022.228/1061245/ws2022228.pdf


by guest
Uncorrected Proof

Water Supply Vol 00 No 0, 3

which usually occurs along the boundaries of hydraulic structures, such as chutes, spillways, and bottom drain pipes. After the
cavitation occurs, such instabilities, including corrosion and erosion, can cause damages to the system. Based on observations
of damages caused by cavitation, the general design guide only dependent on the critical cavitation number is presented in
Table 1.
Aydin & Ozturk (2009) combined and verified CFD numerical method in comparison with experimental methods in aera-
tion of the spillways. They achieved a proper agreement between the two methods. Parsi et al. (2009) constructed a physical
model of the Glaber dam on a scale of 1:30 and studied cavitation in this spillway. The obtained results showed that the cavi-
tation index did not exceed the critical index in any of the studied points in the experiments. Hager (1991) studied the
aeration in uniform flows in chutes. He proved that the average air concentration of the cross-section of the flow is only a
function of the slope of the chute floor. He proposed the following equation:

 90u ¼ 0:75(sin a)0:75


C (1)

where C  90u represents the average uniform air concentration, and α denotes the slope of the chute floor.
Zhang et al. (2013) measured impact pressures of up to 300 MPa due to the collapse and rebound of cavitation bubbles.
Frizell et al. (2013) presented a correlation between the critical cavitation index and the common friction factor for flow
on stepped spillways. Parsaie et al. (2016) simulated cavitation phenomenon along spillway’s flip bucket of the Balaroud
dam using Flow 3D software, and their results indicated that occurrence of cavitation based on cavitation index equal to
0.25 is not possible along the spillway. Kermani et al. (2018) applied fuzzy k-nearest neighbor algorithm to cavitation
damage prediction on dam spillways, and they find that the algorithm was efficient and suitable for this purpose. Ghazi
et al. (2019) simulated the three-dimensional model of Shahid Madani Dam’s spillway using Flow 3D software to study
the probability of occurrence of the cavitation phenomenon. Their results indicated that at any flow rates with a return
period of 1,000 years, the cavitation index is not lower than the critical cavitation number. Barzegari et al. (2019) used a
numerical software to model the flow on the spillway of Aydoghmush Dam, and they find that cavitation did not occur at
any of the considered flow rates. Yusuf & Micovic (2020) studied a prototype-scale modeling of cavitation damage to a
newly resurfaced spillway. Factors that contributed to the cavitation damage for this spillway, including the duration of con-
tinuous spill and the increased cavitation potential of both a smooth concrete surface compared to one that is uniformly
rough and sharp-edged steps compared to steps with rounded edges, were reviewed. Samadi-Boroujeni et al. (2020) focused
on the modeling of the effect of bed roughness height of chute spillways on the cavitation index. Their results indicated that
reducing the roughness height from 2.5 to 1 mm would not change significantly the value of the cavitation index at 95% con-
fidence interval. The focused of the present study is on the analysis of the effects of convergence angle of sidewalls on the
cavitation index of the arc-plan ogee spillways. Three important limitations for the selection of a spillway location are as
follows: (1) narrow width of the channel or reservoir outlet; (2) narrow width of the downstream river or narrow diameter
of downstream exiting tunnels; and (3) the high amount of kinetic energy of the flow in the spillway downstream, which
cause cavitation phenomenon and impose exorbitant costs of the stilling basin structure. The considered ogee spillway
can fade all of three aforementioned limitations. The narrow width of the channel or reservoir outlet can eliminate by increas-
ing length of spillway by an arc plan. The narrow width of the downstream river can consider by convergence angle of
sidewalls. The cavitation phenomenon can control by stepped structure at downstream of the proposed spillway. To this
end, a physical model of such ogee spillway was constructed on the 1:50 scale. Then, it was tested with four different

Table 1 | General design guidelines dependent on the critical cavitation number

Design consideration Cavitation index s

Needs no protection against cavitation . 1:80


The surface level can be protected by surface modification (modifying surface irregularities) 0.25–1.80
Design a modification (for example, increasing the curvature of the boundary) 0.25–017
Protection using air galleries 0.12–0.17
The surface cannot be protected; as a result, the new design is needed , 0:12

Downloaded from http://iwaponline.com/ws/article-pdf/doi/10.2166/ws.2022.228/1061245/ws2022228.pdf


by guest
Uncorrected Proof

Water Supply Vol 00 No 0, 4

convergence angles, including 0°, 60°, 90°, and 120° relative to the spillway’s sidewalls. Each one of the convergent cases was
examined at six different flow rates per unit width, including 6.74, 16.88, 27.02, 30.39, 40.52, and 48.42 (l/s)/m.

MATERIAL AND METHODS


The experiments were conducted at the Soil Conservation and Watershed Management Research Institute (SCWMRI) in
Tehran, Iran. The structure of the spillway model was constructed using polyethylene waterproof material. Plexiglas was
used to construct the walls and bottom of the channel prototype. Table 2 summarized the design parameters for 3D conver-
ging ogee spillway models with a 1:50 scale and the design parameters of the channel prototype.
As it is well-known the impact of gravity was typically more important compared to the impact of viscosity and surface
tension for free surface flows as considered spillways in the present study (Falvey 1990; Crowe et al. 2009). Therefore, the
model and prototype were generally based on the Froude (Fr) similarity in the scale relationship without considering the vis-
cosity effect. This occurs when the Reynolds number is larger than a certain limit (¼104).
The experiment’s first phase was executed at an angle of 120° with Lch/L ¼ 0.21. It should be stated that L is the crest length
and Lch denotes the downstream channel width. For all experiments, the discharge was monitored and measured using a
sharp triangular weir with an apex angle of 90° at the channel output, where the channel was kept at a rough zero-slope.
The water surface profiles were monitored using a point gage along the centerline of the spillway with the other two specific
lines located on either side. Uncertainties associated with experimental measurement for water elevation reading were
+1 mm. To monitor the pressure on the spillway, several taps were placed along the centerline of the spillway as well as
the two abovementioned specific lines which included 12 stations with known coordinates and then they were connected
to the piezometers board. This approach to measure pressure was also used by Johnson and Savage (2006) and Naghavi
et al. (2011). The values of the velocity were calculated by using the continuity equation with the known flow rate, width
and flow depth in each station. Given the fixed downstream channel width, creating each angle changes the crest length
of the spillway model due to operational constraints in the design of the prototype. Figure 1 illustrates the model from
four angles, and Figure 2 presents a side view of the model.
In the second phase of the experiments, the constructed physical model was tested with three convergence angles between
the sidewalls, including 0°, 60°, and 90° with three different Lch/L ratios of 0.98, 0.32, and 0.26. In order to consider the con-
stant downstream channel width, a changing spillway’s length of the crest for every angel θ was considered. As it can be find
in Table 3, flow rates were considered with variation of θ based on the unit discharge flow rate (q) at the crest in order to
maintain constant conditions for all of the convergence angles of different tests.
2:5
The flow rate of the model with a scale of 1:50 was calculated using equation Qm =Qp ¼ Dm =Dp , in which Qm and Qp
respectively represent the flow rate in the model and the prototype, while Dm and Dp respectively denote the longitudinal
dimension of the model and the prototype. Since the flow rate per unit width in all angles had to be identical to that in
the model with an angle of 120°, each of the presented flow rates was multiplied by the ratio of the crest length at the desired
angle to the crest length at the angle of 120°. The velocity-dependent index is typically used to control the cavitation phenom-
enon. Given the particular shape of the ogee spillway of the current study, it was essential to calculate this index and the
probability of the occurrence of the cavitation. This method aims to achieve an index higher than the critical cavitation

Table 2 | Model parameters and prototype design dimensions

Design elements Prototype dimensions Model dimensions

Convergence angles 120° 120° 90° 60° 0°


Crest length (m) 42.83 0.837 0.712 0.578 0.183
3 1
Design discharge (m s ) 398 0.0225 0.0182 0.0158 0.0048
Maximum discharge (m3 s1 ) 717 0.0405 0.0344 0.028 0.00886
Maximum Head (m) 5 0.1
Spillway height (m) 7.8 0.156
Design head (m) 3 0.06
Downstream channel width (m) 9 0.18

Downloaded from http://iwaponline.com/ws/article-pdf/doi/10.2166/ws.2022.228/1061245/ws2022228.pdf


by guest
Uncorrected Proof

Water Supply Vol 00 No 0, 5

Figure 1 | The plan of the model at each of the four angles: (a) 0°, (b) 60°, (c) 90°, (d) 120°, (e) a comparison of the views of the model at the
four angles.

Figure 2 | A cross-section of the model.

Table 3 | Characteristics of the flow rates both prototype and model at different angles

Lch Lch Lch Lch


Discharge (prototype) (m3/s) 0° ¼ 0:98 60° ¼ 0:32 90° ¼ 0:26 120° ¼ 0:21 q L/s/m
L L L L

100 1.24 3.9 4.8 5.65 6.74


250 3.09 9.76 11.99 14.14 16.88
400 4.59 15.61 19.18 22.62 27.02
450 5.56 17.56 21.58 25.46 30.39
600 7.42 23.41 28.77 33.94 40.52
717 8.86 27.98 34.39 40.56 48.42

Downloaded from http://iwaponline.com/ws/article-pdf/doi/10.2166/ws.2022.228/1061245/ws2022228.pdf


by guest
Uncorrected Proof

Water Supply Vol 00 No 0, 6

index, which was provided in Table 1. In order to estimate this index, the following expression si . s was used:

Po  Pv
s¼ (2)
r V 2 =2

Based on the results obtained by Falovi (1990), the critical cavitation index can be calculated using the following equation
for a single roughness:

h  hv
si ¼ (3)
V02 =2g

where h denotes the height of absolute pressure, hv indicates the height of the vapor pressure, and v0 represents the velocity
near the surface. Arendt (1979) showed that cavitation could occur on a surface with uniform roughness. Then, he presented
the equation below using Darcy – Weisbach ƒ friction factor.

si ¼ 4f (4)

Folovi (1990) presented a complete set for another rough. For example, he showed that the tunnel spillways with indexes
equal to or lower than 0.2 are influenced by cavitation. However, tunnel spillways with indexes higher than 0.2 are not influ-
enced by cavitation.

RESULTS AND DISCUSSION


In the present study, the physical model of the spillway of Garmi Chai dam in East Azerbaijan province, Iran with a scale of
1:50 was used. This spillway has an angle of 120 degrees and it’s Qpmf (probable maximum flood) and maximum allowable
head respectively are 717 m3/s and 5 m. The spillway tested in the first part of the experiments at an angle of 120 degrees with
Lch/L ¼ 0.21 at the highest allowable head and it could not pass the maximum flow rate. The reason for this situation was the
sharp angle and excessive shrinkage of the ratio of downstream channel width to spillway crest length. Therefore, the exper-
iments in the second phase led to the use of smaller angles that had a larger Lch/L ratio (i.e. increasing the width of the
downstream channel). By considering this issue, 90, 60 and 0 degree angles were selected and tested in the experiments.
Based on the obtained experimental data, a continuous and supercritical flow ran over the studied spillway at different
angles can be observed. This occurred in all cases except the following conditions: at the angles of 90° and 120° with
Lch/L ratios of 0.26 and 0.21, where the spillway was submerged in flow rates per unit width of 48.42 and 40.52 (l/s)/m,
respectively. Moreover, the flow regime changed to subcritical flow. The convergence of the sidewalls at 60°, 90° and 120°
led to the occurrence of the rooster tail phenomenon at the toe of the spillway, as shown in Figure 3.
According to the experimental observations, the flows over two sides of the spillway were similar, which agrees with the
results reported by Frudi et al. (2012). Therefore, the average values of the cavitation index in removable sectors were com-
pared at different flow rates and angles. The cavitation index for the angle of 120° and flow rate per unit width of 16.88 (l/s)/m

Figure 3 | Rooster tail phenomenon.

Downloaded from http://iwaponline.com/ws/article-pdf/doi/10.2166/ws.2022.228/1061245/ws2022228.pdf


by guest
Uncorrected Proof

Water Supply Vol 00 No 0, 7

Figure 4 | The cavitation index along the longitudinal direction of a dimensionless spillway with head design in the flow rate per unit width of
16.88 (l/s/m).

in the five different sectors presented in Figure 4 to provide a better understanding of the similarity of the flows over both sides
of the spillway. The lowest cavitation index in all cases was observed at the station with an X/Hd of 2.42. This parameter was
found to be 1.54 at the angle of 0° with Lch/L of 0.98, while being equal to 2.20 at the angle of 60° with Lch/L of 0.32, at the
constant flow rate per unit width of 40.5 (l/s)/m.
The cavitation index was calculated at 2.18 for the angle of 90°, Lch/L ¼ 0.26, and flow rate per unit width of 30.39 (l/s)/m,
while it was found to be 1.95 for the angle of 120°, Lch/L ¼ 0.21, and flow rate per unit width of 16.88 (l/s)/m. As can be seen,
the most critical index is not necessarily observed in the maximum flow rate as previously discussed by USBR (1995). How-
ever, the obtained results indicated that with the rise in the flow rate, the cavitation index decreases relatively at both chute
and crest of the spillway, whereas increasing at its toe. Figure 5(a)–5(d) depicts the cavitation index for different flow rates per
width at various angles, including 0°, 60°, 90°, and 120°. The vertical axis of the chart was plotted for only the cavitation
indexes of up to 15 since the purpose of plotting the charts is to show the changes in the values closer to the critical indexes.
As can be seen in Figure 5, the areas prone to damages caused by cavitation are the bottom chute and toe of the spillway,
given the minimal values of the cavitation index at different angles based on the critical cavitation index (Cr ¼ 0.2σ).
The variations of the cavitation index for four convergence angles 0°, 60°, 90°, and 120° were presented in Figure 6. As it can
be seen in the area between the crest and the middle of the chute of the spillway, changes in the convergence angle have no
effect on the cavitation index. However, at the end of the chute and the toe of the spillway, the cavitation index rises with the
convergence angle. The Froude number over the spillway for four convergence angles 0°, 60°, 90°, and 120° were depicted in
Figure 7. As it can be seen the convergence angle for a fixed flow rate per unit width increases when the Froude number
decreases. These results indicate that if the convergence angle increases, the gravity force surpasses the inertial force in
the hydraulic flow. Consequently, the growth rate of the depth increases with respect to the velocity.

SUMMARY AND CONCLUSION


The present study focused on the experimental study of the convergence angle effects on the convergence angle of an ogee
spillway’s sidewalls with an arc in the plan. The convergence angle was varied in the range of 0° to 120°. Moreover, six differ-
ent flow rates per unit width were considered in the range of 6.74–48.42 (l/s)/m. According to experimental observations and
analysis of them, the similarity of the flow at both sides of the spillway was found for all considered convergence angles. The
lowest cavitation index was 1.54 at an angle of 0° with X/Hd ¼ 2.33, Lch/L ¼ 0.98, and flow rate per unit width of 40.52 (l/s)/m.
It was observed that the most critical index was not necessarily observed at the maximum flow rate. As the flow rate
increased, the cavitation index relatively declined at both crest and chute of the spillway while growing at its toe. Such
changes in the cavitation index were observed for all convergence angles 0°, 60°, 90°, and 120°. In the area between the
crest in the middle of the chute and the toe of the spillway, the cavitation index increased with the rise of the convergence

Downloaded from http://iwaponline.com/ws/article-pdf/doi/10.2166/ws.2022.228/1061245/ws2022228.pdf


by guest
Uncorrected Proof

Water Supply Vol 00 No 0, 8

Figure 5 | Changes in the cavitation index for different flow rates per unit width at the angles of (a) 0°, (b) 60°, (c) 90°, (d) 120°.

Figure 6 | Changes in the cavitation index at different convergence angles in the fixed flow rate per unit width of 27.02 (l/s)/m.

Downloaded from http://iwaponline.com/ws/article-pdf/doi/10.2166/ws.2022.228/1061245/ws2022228.pdf


by guest
Uncorrected Proof

Water Supply Vol 00 No 0, 9

Figure 7 | The Froude number over the spillway at convergence different angles.

angle and the consequent reduction in the ratio of Lch/L. For considering more sets of experiments, the use of numerical
simulation tools can be considered as numerical experiments in the future research. The presented results of physical exper-
iments can be used to validate numerical model. Finally it can be said that the proposed ogee spillway can be used by design
engineers with execution details for real field projects.

DISCLOSURE STATEMENT
The authors declare that they have no known competing financial interests or personal relationships that could have
appeared to influence the work reported in this paper.

DATA AVAILABILITY STATEMENT


All relevant data are included in the paper or its Supplementary Information.

CONFLICTS OF INTEREST STATEMENT


The authors declare there is no conflict.

REFERENCES

Adama Maiga, M., Coutier-Delgosha, O. & Buisine, D. 2016 Cavitation in a hydraulic system: the influence of the distributor geometry on
cavitation inception and study of the interactions between bubbles. International Journal of Engine Research 17 (5), 543–555.
Arami-Fadafan, M., Hessami-Kermani, M. R. & Barati, R. 2018a Applications of improved moving particle semi-implicit method for
numerical simulation of flow over hydraulic structures. In: progress in River Engineering & Hydraulic Structures. International energy
and environment foundation, pp. 55–92, chapter four.
Arami-Fadafan, M., Moghaddam, M. A. & Barati, R. 2018b Reduce spillways cavitation damages through accurate prediction of bottom air
concentration. In: Progress in River Engineering & Hydraulic Structures. International Energy and Environment Foundation, pp. 33–54,
Chapter Three.
Aydin, M. C. & Ozturk, M. 2009 Verification and validation of a computational fluid dynamics (CFD) model for air entrainment at spillway
aerators. Canadian Journal of Civil Engineering 36 (5), 826–836.
Ball, J. W. 1976 Cavitation from surface irregularities in high velocity. Journal of the Hydraulics Division 102 (9), 1283–1297.
Bananmah, M., Nikoo, M. R., Nematollahi, B. & Sadegh, M. 2020 Optimizing chute-flip bucket system based on meta-modelling approach.
Canadian Journal of Civil Engineering 47 (5), 584–595.

Downloaded from http://iwaponline.com/ws/article-pdf/doi/10.2166/ws.2022.228/1061245/ws2022228.pdf


by guest
Uncorrected Proof

Water Supply Vol 00 No 0, 10

Barati, R. 2012 Discussion of ‘Parameter estimation of the nonlinear Muskingum model using parameter-setting-free harmony search’ by
Zong Woo Geem. Journal of Hydrologic Engineering 17 (12), 1414–1416.
Barzegari, M., Sobhkhiz Foumani, R., Isari, M., Tarinejad, R. & Alavi, S. A. 2019 Numerical investigation of cavitation on spillways. A case
study: Aydoghmush dam. International Journal of Numerical Methods in Civil Engineering 4 (1), 1–9.
Chanson, H. 1988 Study of Air Entrainment and Aeration Devices on Spillway Model. PHD Thesis, University of Canterbury, Christchurch,
New Zealand, p. 0011.
Crowe, C. T., Elger, D. F., Williams, B. & Roberson, J. 2009 Engineering Fluid Mechanics, Vol. 714. Wiley, XI, New York, p. 36.
Falvey, H. T. 1983 Prevention of cavitation on chutes and spillways. Frontiers in Hydraulic engineering, 432–437.
Falvey, H. T. 1990 Cavitation in Chutes and Spillways. US Department of the Interior, Bureau of Reclamation, Denver, p. 145.
Frizell, K. W., Renna, F. M. & Matos, J. 2013 Cavitation potential of flow on stepped spillways. Journal of Hydraulic Engineering 139 (6),
630–636.
Ghazi, B., Daneshfaraz, R. & Jeihouni, E. 2019 Numerical investigation of hydraulic characteristics and prediction of cavitation number in
Shahid Madani Dam’s Spillway. Journal of Groundwater Science and Engineering 7 (4), 323–332.
Hager, W. H. 1991 Uniform aerated chute flow. Journal of Hydraulic Engineering 117 (4), 528–533.
Hosseini, K., Nodoushan, E. J., Barati, R. & Shahheydari, H. 2016 Optimal design of labyrinth spillways using meta-heuristic algorithms.
KSCE Journal of Civil Engineering 20 (1), 468–477.
Karimi Pirmoosaei, F. & Mardookhpour, A. 2020 Numerical simulation of cavitation phenomenon in the stepped spillway with fluent
software. Civil Infrastructure Researches 6 (1), 127–140.
Kermani, E. F., Barani, G. A. & Ghaeini-Hessaroeyeh, M. 2013 Investigation of cavitation damage levels on spillways. World Applied Sciences
Journal 21 (1), 73–78.
Kermani, E. F., Barani, G. A. & Hessaroeyeh, M. G. 2018 Cavitation damage prediction on dam spillways using fuzzy-KNN modeling. Journal
of Applied Fluid Mechanics 11 (2), 323–329.
Kocaer, Ö. & Yarar, A. 2020 Experimental and numerical investigation of flow over ogee spillway. Water Resources Management 34 (13),
3949–3965.
Koen, J. 2017 Artificial Aeration on Stepped Spillways with Piers and Flares to Mitigate Cavitation Damage. Doctoral dissertation,
Stellenbosch University, Stellenbosch.
Kramer, K. 2004 Development of Aerated Chute Flow. PHD Thesis, Mitteilungen010, Versuchsanstalt für Wasserbau Hydrologie und
Glaziologie der Eidgenössischen, Technischen Hochschule Zürich, p. 2116.
Parsaie, A., Dehdar-Behbahani, S. & Haghiabi, A. H. 2016 Numerical modeling of cavitation on spillway’s flip bucket. Frontiers of Structural
and Civil Engineering 10 (4), 438–444.
Parsi, E., Behravandi, M., Fathi Moghadam, M. & Kazem zadeh, A. 2009 Study of cavitation phenomena over the spillway using physical
models. In 1th Conference of Iranian Hydraulic. p. 2110
Rafi, M., Ali, A., Qadir, G. & Ali, R. 2012 Modeling the Mangla dam spillway for cavitation and aerators optimization. Journal of Water
Resource and Protection 4 (12), 1051.
Rajasekhar, P. Y. V. G., Santhosh, Y. V. G. & Soma Sekhar, S. 2014 Physical and numerical model studies on cavitation phenomenon – a
study on Nagarjuna Sagar spillway. International Journal of Recent Development in Engineering and Technology 2 (1), 1–10.
Samadi-Boroujeni, H., Abbasi, S., Altaee, A. & Fattahi-Nafchi, R. 2020 Numerical and physical modeling of the effect of roughness height on
cavitation index in chute spillways. International Journal of Civil Engineering 18 (5), 539–550.
Shahheydari, H., Nodoshan, E. J., Barati, R. & Moghadam, M. A. 2015 Discharge coefficient and energy dissipation over stepped spillway
under skimming flow regime. KSCE Journal of Civil Engineering 19 (4), 1174–1182.
Tajnesaie, M., Jafari Nodoushan, E., Barati, R. & Azhdary Moghadam, M. 2020 Performance comparison of four turbulence models for
modeling of secondary flow cells in simple trapezoidal channels. ISH Journal of Hydraulic Engineering 26 (2), 187–197.
United States department of the Interior, Bureau of reclamation (USBR), Hydraulic techniques, U.S. Government printing office, Denver, 0011.
Usta, E. 2014 Numerical Investigation of Hydraulic Characteristics of Laleli Dam Spillway and Comparison with Physical Model Study.
Master’s Thesis.
Wei, J. I. 2006 Summary on cavitation research of ogee-section in spillway tunnels. Journal of Heilongjiang Hydraulic Engineering College 1.
Yusuf, F. & Micovic, Z. 2020 Prototype-scale investigation of spillway cavitation damage and numerical modeling of mitigation options.
Journal of Hydraulic Engineering 146 (2), 04019057.
Zhang, H., Han, B., Yu, X. G. & Ju, D. Y. 2013 Numerical and experimental studies of cavitation behavior in water-jet cavitation peening
processing. Shock and Vibration 20 (5), 895–905.

First received 25 January 2022; accepted in revised form 3 June 2022. Available online 9 June 2022

Downloaded fromView
http://iwaponline.com/ws/article-pdf/doi/10.2166/ws.2022.228/1061245/ws2022228.pdf
publication stats
by guest

You might also like